29
CHEMICAL AGENTS AND RELATED OCCUPATIONS volume 100 F A review oF humAn cArcinogens This publication represents the views and expert opinions of an IARC Working Group on the Evaluation of Carcinogenic Risks to Humans, which met in Lyon, 20-27 October 2009 LYON, FRANCE - 2012 iArc monogrAphs on the evAluAtion oF cArcinogenic risks to humAns

chemical agents and related occupations · Vinyl chloride is used primarily (> 95%) in the manufacture of polyvinyl chloride (PVC), which comprises about 12% of the total use of plastic

  • Upload
    others

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

  • chemical agents and related occupations

    volume 100 FA review oF humAn cArcinogens

    this publication represents the views and expertopinions of an iarc Working group on the

    evaluation of carcinogenic risks to humans, which met in lyon, 20-27 october 2009

    lyon, france - 2012

    iArc monogrAphs on the evAluAtion

    oF cArcinogenic risks to humAns

  • VINYL CHLORIDE Vinyl chloride was considered by previous IARC Working Groups in 1974, 1978, 1987, and 2007 (IARC, 1974, 1979, 1987, 2008). Since that time new data have become available, which have been incorporated in this Monograph, and taken into consideration in the present evaluation.

    1. Exposure Data

    1.1 Identification of the agent

    From IARC (2008) and Lide (2008) Chem. Abstr. Serv. Reg. No.: 75-01-4 Chem. Abstr. Serv. Name: Chloroethene

    H2C CH Cl

    C2H3C1 Relative molecular mass: 62.5 Description: Colourless gas, with a mild, sweet odour Boiling-point: -13.4 to -13.8 °C Solubility: Slightly soluble in water (1.1 g/L at 25 °C); soluble in ethanol; very soluble in diethyl ether, carbon tetrachloride and benzene Conversion factor: 1 ppm = 2.6 mg/m3

    1.2 Uses

    Vinyl chloride is used primarily (> 95%) in the manufacture of polyvinyl chloride (PVC), which comprises about 12% of the total use of plastic worldwide (WHO, 1999). The largest use of PVC

    is in the production of plastic piping. Other important uses are in floor coverings, consumer goods, electrical applications and in the transport sector. About 1% of PVC is used to produce vinyl chloride/vinyl acetate copolymer. Minor uses of vinyl chloride (monomer) include the manufacture of chlorinated solvents (primarily 10000 tonnes per year of 1,1,1-trichloroethane) and the production of ethylene diamine for the manufacture of resins (WHO, 1999; European Commission, 2003).

    Vinyl chloride has been used in the past as a refrigerant, as an extraction solvent for heat-sensitive materials, in the production of chloroacetaldehyde, as an aerosol propellant and in drugs and cosmetic products; these uses were banned in the United States of America (USA) by the Environmental Protection Agency in 1974 (IARC, 2008).

    1.3 Human exposure

    1.3.1 Occupational exposure

    The main route of occupational exposure to vinyl chloride is by inhalation, which occurs primarily in vinyl chloride/PVC plants and in PVC-processing plants. Only few exposure

    451

  •  

    IARC MONOGRAPHS – 100F

    measurements have been reported, but estimates from the chemical industry indicate that exposure to vinyl chloride monomer (VCM) amounted to several thousands of milligrams per cubic metre in the 1940s and 1950s, and were several hundreds of milligrams per cubic metre in the 1960s and early 1970s. After its recognition as a human carcinogen (IARC, 1974), occupational exposure standards were set at approximately 13–26 mg/m3 [5–10 ppm] in most countries in the 1970s (Fleig & Thiess, 1974; NTP, 2005; IARC, 2008).

    CAREX (CARcinogen EXposure) is an international information system on occupational exposure to known and suspected carcinogens based on data collected in the European Union (EU) from 1990 to 1993. The CAREX database provides selected exposure data and documented estimates of the number of exposed workers by country, carcinogen, and industry (Kauppinen et al., 2000). Table 1.1 presents the estimated numbers of workers exposed to vinyl chloride in the EU for the top-10 industries (CAREX, 1999).

    From the US National Occupational Exposure Survey (1981–83) it was estimated that approximately 81 300 workers (including approximately 28 400 women) were potentially exposed to vinyl chloride (NIOSH, 1990).

    A report from the Centers for Disease Control and Prevention (CDC) in the USA concluded that the development and acceptance by the PVC-manufacturing industry of a closed-loop polymerization process in the late 1970s “almost completely eliminated worker exposures” (CDC, 1997). Even after the late 1970s, however, high concentrations may still be encountered and were in fact reported in some countries (IARC, 2008).

    (a) Production of vinyl chloride and its derivatives

    In vinyl-chloride production, workers may be exposed to ethylene dichloride and to catalysts such as iron(III)chloride. In PVC production,

    concurrent exposure to PVC-dust may occur (Casula et al., 1977).

    Measurements of VCM concentrations in indoor air of vinyl chloride/PVC production plants have been summarized in IARC Monograph Volume 97 (IARC, 2008). In a study on occupational exposure to vinyl chloride that was not included in the previous Monograph, Zhu et al. (2005) reported the exposure to VCM of workers in a plant in the People’s Republic of China. Concentrations in air of VCM at different worksites in the plant ranged from 0.3 to 17.8 ppm [0.8–48.4 mg/m3]; the geometric mean concentration was 2.6 ppm [7.1 mg/m3].

    (b) PVC processing

    Measured concentrations of VCM in PVC-processing plants were considerably lower than those in plants where vinyl chloride and PVC were produced (IARC, 2008). Improvements in PVC production in the 1970s resulted in a much lower content of residual VCM in PVC resin. The lower monomer content led automatically to reduced concentrations of vinyl chloride in the workplace air of PVC-processing factories, reaching values of

  • Vinyl chloride

    Table 1.1 Estimated numbers of workers exposed to vinyl chloride in the European Union (top-10 industries)

    Industry, occupational activity

    Manufacture of plastic products, not elsewhere classified 9100 Manufacture of other chemical products 7600 Manufacture of fabricated metal products, except machinery and equipment 2900 Manufacture of machinery, except electrical 2400 Services allied to transport 1300 Manufacture of electrical machinery, apparatus, appliances 980 Education services 870 Construction 800 Petroleum refineries 570 TOTAL 39600

    Manufacture of industrial chemicals 10400

    From CAREX (1999)

    1.3.2 Non-occupational exposure

    The general population is potentially exposed to vinyl chloride through inhalation of contaminated air, ingestion of contaminated drinking-water and foods, or dermal contact with consumer products. However, the exposure levels for the majority of the population are very low (NTP, 2005).

    (a) Ambient air

    Vinyl chloride is released into the environment in emissions and effluents from the plastics industry. Atmospheric concentrations of VCM in ambient air are low (usually

  • IARC MONOGRAPHS – 100F

    (c) Residues in PVC resin and products

    PVC products may contain VCM as a residue from production and release it in the air. In a German survey (1976–77), the following articles released VCM at levels > 0.05 ppm [0.13 mg/m3] by off-gassing in the air: bathroom tiles, piping, plastic bottles for table oil, and kitchen wrapping-film. The highest concentrations were observed to come from vinyl music records, with values of 20–50 ppm measured for nine of 14 records sampled, but even higher in some of the others. The VCM concentrations released by toys, kitchen utensils, food wrappings, wall-paper, and car interiors were

  • Vinyl chloride

    were required to have been exposed to VCM for at least one year before 31 December 1972 and to have been employed in or after 1942. A second major update of this cohort was published by Wong et al. (1991). A third major follow-up included 10 109 subjects and provided an update of the vital status through to 31 December 1995 (Mundt et al., 2000).

    The European cohort study was conducted in four countries (Italy, Norway, Sweden and the United Kingdom). It included workers from 19 factories: 11 of these produced VCM/PVC, two produced VCM only, five produced PVC only and one was a PVC-processing plant. Male workers who had been employed for at least one year in 1942–1972 in jobs that entailed exposure to VCM were included (Simonato et al., 1991). An update of the study (Ward et al., 2001) analysed incidence and mortality through to the latest year for which data were available in each country, which ranged between 1993 and 1997.

    2.1 Angiosarcoma of the liver

    In both multicentre cohort studies (Mundt et al., 2000; Ward et al., 2001) a substantial excess of ASL in exposed workers was found (see Table  2.1 available at http://monographs. iarc.fr/ENG/Monographs/vol100F/100F-26 Table2.1.pdf). This tumour is extremely rare in the general population and it is not possible to calculate an SMR or SIR, because age- and calendar time-specific reference rates are not available. In the study from the US, 33 of the 80 deaths from cancer of the liver and biliary tract were identified from the death certificate as due to ASL. A total of 48 deaths due to ASL were identified by combining information from death certificates with that from a registry of ASL-cases that were related to exposure to VCM. This registry is maintained and updated by the Association of Plastics Manufacturers of Europe.

    In the European study there were 53 deaths from primary liver cancer and 18 incident cases

    of liver cancer. This total of 71 cases comprised 37 ASL, 10 HCC, 7 cases of other known histology, and 17 cases of an unspecified type of liver cancer. [The Working Group noted that the authors searched for the best evidence for diagnosis of liver cancers by reviewing all available documentation, including death certificates, cancer-registry records, medical records, and listings of ASL from two registries.]

    In both studies, the risk for ASL increased strongly with duration of exposure to vinyl chloride. In the European study, there was also a clear trend of higher risk with increasing cumulative exposure. Multiple cases of ASL were also reported in one smaller cohort study (Thériault & Allard, 1981). Two cases of ASL were reported among hairdressers and barbers who had been exposed to vinyl chloride for 4–5-year periods in the late 1960s and early 1970s, when it was used as a propellant in hairspray (Infante et al., 2009).

    2.2 Hepatocellular carcinoma

    The assessment of vinyl chloride as a cause of HCC is complicated because many studies do not have histological or other definitive clinical information to discriminate HCC from ASL and/or secondary neoplasms (see Table 2.1 online). In the US multicentre study, mortality from cancers of the liver and biliary tract (ICD9code, 155–156) was increased (SMR 3.6, 95%CI: 2.8–4.5; 80 deaths). Of the 80 deaths, 48 were identified as ASL. The diagnosis of HCC among the remaining deaths was not verified.

    In an internal analysis of the European multicentre cohort (Ward et al., 2001) based on 10 verified cases of HCC, the risk increased significantly and substantially with duration of employment and with cumulative exposure to vinyl chloride. The relative risk for workers with the longest duration of employment (> 26 years) was 35 (95%CI: 3.3–377) compared with workers with

  • IARC MONOGRAPHS – 100F

    was included in the European study – indicated 12 confirmed cases of HCC (Pirastu et al., 2003). The maximal overlap between these two analyses was four cases, since only four HCC from Italy were included in the multicentre cohort. In this subcohort, the incidence of HCC again increased significantly with cumulative exposure to vinyl chloride. There was suggestive evidence that the risk for HCC from vinyl chloride is substantially higher among workers who are infected with hepatitis B virus (Wong et al., 2003), or who report high levels of alcoholic beverage consumption (Mastrangelo et al., 2004).

    A meta-analysis of cohort studies of vinyl choride-exposed workers published up to 2002 (Boffetta et al., 2003) was based on eight independent studies, i.e. two multicentric investigations (Mundt et al., 2000; Ward et al., 2001) and six additional, smaller studies (Thériault & Allard, 1981; Weber et al., 1981; Smulevich et al., 1988; Laplanche et al., 1992; Huang, 1996; Wong et al., 2002) (p-value for the test for heterogeneity was ≥  0.01). Six of these eight studies reported results for liver cancer, but these were considered to be too heterogeneous to be included in a meta-analysis because for ‘liver cancer overall’ and for ‘liver cancer other than ASL’, the p-value for heterogeneity was  20 years) and for those first employed before 1960. Four of the 12 observed deaths were from angiosarcomas for which the site was unknown. The increased mortality from neoplasms of connective and soft tissue persisted even after exclusion of these four angiosarcomas. [This presumes that the malignant neoplasms of connective and soft tissue were mis-classified deaths from angiosarcoma of the liver.]

    The findings mentioned above were not supported by results from the European multicentre study, in which the number of deaths from connective-tissue neoplasms was too small

    456

    http://monographs.iarc.fr/ENG/Monographs/vol100F/100F-26-Table2.2.pdfhttp://monographs.iarc.fr/ENG/Monographs/vol100F/100F-26-Table2.2.pdfhttp:0.84�1.83http:1.15�2.62http:1.04�4.39http:2.00�4.39

  •  

    Vinyl chloride

    for an evaluation of exposure–response (Ward et al., 2001): there were six observed deaths from neoplasms of connective and soft tissue (SMR = 1.9, 95%CI: 0.7–4.1), but in a re-evaluation of the diagnoses three of the six deaths coded as tumours of the connective tissue were found to be ASL. [The Working Group noted that, although a statistically significant increase in mortality from neoplasms of connective and soft tissue was found in the US study, the discrepant results with the European study and the difficulties in arriving at a correct diagnosis and coding of the tumour site for this type of neoplasm, complicate an evaluation of these findings.]

    2.5 Other cancers

    The Working Group did not find strong evidence for associations of exposure to vinyl chloride with cancers of the brain or the lymphatic and haematopoeitic tissues, with melanoma of the skin (see Table 2.3 available at http://monographs.iarc.fr/ENG/Monographs/ vol100F/100F-26-Table2.3.pdf, Table 2.4 available at http://monographs.iarc.fr/ENG/Monographs/ vol100F/100F-26-Table2.4.pdf, and Table  2.5 available at http://monographs.iarc.fr/ENG/ Monographs/vol100F/100F-26-Table2.5.pdf ). Although the associations found for these cancers in specific studies may reflect true increases in risk, the findings were inconsistent between studies, no clear exposure–response relationships were found in the European multicentre study (Ward et al., 2001), and, for several of the sites, the numbers of observed/expected cases were small.

    No conclusion could be reached for breast cancer since the available studies included too few women.

    2.6 Synthesis

    There is compelling evidence that exposure to vinyl chloride is associated with angiosarcoma of the liver, and strong evidence that it is associated with hepatocellular carcinoma. Together with the observation that vinyl chloride increases the risk for liver cirrhosis, which is a known risk factor for hepatocellular carcinoma, the findings from two large multicentre cohort studies provide convincing evidence that vinyl chloride causes hepatocellular carcinoma as well as angiosarcoma of the liver. There is contradictory evidence that exposure to vinyl chloride is associated with malignant neoplasms of connective and soft tissue, and inconsistent or scanty evidence that it is associated with cancers of the lung, brain, lymphohaematopoietic system, and breast, or with melanoma of the skin.

    3. Cancer in Experimental Animals

    The carcinogenicity of vinyl chloride has been studied intensively and repeatedly in experimental animals, with a wide range of concentrations, spanning orders of magnitude. The many studies consistently showed hepatic and extra-hepatic angiosarcomas in mice and rats. Various other malignant neoplasms also occurred at several anatomical sites. However, the reporting of the results has often been incomplete, and the outcomes of many studies are available only from summary tables in the published literature, in which technical details are given in footnotes.

    Studies of the carcinogenicity of vinyl chloride in experimental animals after oral administration, inhalation, subcutaneous injection, intraperitoneal injection, and transplacental and perinatal exposure have been reviewed in previous IARC Monographs (IARC, 1974, 1979, 1987, 2008). No studies have been published since the most recent evaluation (IARC, 2008). The

    457

    http://monographs.iarc.fr/ENG/Monographs/vol100F/100F-26-Table2.3.pdfhttp://monographs.iarc.fr/ENG/Monographs/vol100F/100F-26-Table2.3.pdfhttp://monographs.iarc.fr/ENG/Monographs/vol100F/100F-26-Table2.4.pdfhttp://monographs.iarc.fr/ENG/Monographs/vol100F/100F-26-Table2.4.pdfhttp://monographs.iarc.fr/ENG/Monographs/vol100F/100F-26-Table2.5.pdfhttp://monographs.iarc.fr/ENG/Monographs/vol100F/100F-26-Table2.5.pdf

  • IARC MONOGRAPHS – 100F

    following is a summary of the available data (see also Table 3.1).

    3.1 Inhalation exposure

    Vinyl chloride was tested by inhalation exposure in several studies in mice (Holmberg et al., 1976; Lee et al., 1978; Hong et al., 1981; Maltoni et al., 1981; Drew et al., 1983; Suzuki, 1983), in several studies in rats (Lee et al., 1978; Feron et al., 1979; Feron & Kroes, 1979; Groth et al., 1981; Kurliandskiĭ et al., 1981; Maltoni et al., 1981; Drew et al., 1983), and in two studies in hamsters (Maltoni et al., 1981; Drew et al., 1983). Male and female animals of all three species were included, although some experiments were carried out only in one sex. Vinyl chloride induced hepatic angiosarcomas in three experiments in mice and in eight experiments in rats; a dose–response was observed for hepatic angiosarcomas in both species over a wide range of exposures. Extrahepatic angiosarcomas related to treatment with vinyl chloride were observed in three studies in mice and one study in rats. Vinyl chloride increased the incidence of malignant mammary tumours in seven experiments in mice, in two experiments in one study in rats, and in one study in hamsters. Exposure to vinyl chloride increased the incidence of skin epitheliomas in one study in rats and one study in hamsters, and of skin carcinomas in another study in hamsters. It increased the incidence of Zymbal gland carcinomas in three experiments in rats, with a dose–response pattern in one experiment. In mice, vinyl chloride increased the incidence of benign lung tumours in six experiments, and of lung carcinomas in two experiments. It also increased the incidence of nasal cavity carcinomas in one study in rats, of hepatocellular carcinomas in two experiments in rats, of glandular adenomas in one study in hamsters, and of benign fore-stomach tumours in another study in hamsters.

    In one study in rats, combined oral administration of ethanol and inhalation exposure to vinyl chloride increased the incidence of hepatic angiosarcomas compared with exposure to vinyl chloride alone (Radike et al., 1981).

    3.2 Oral administration

    Vinyl chloride was tested by oral administration in four experiments in male and female rats (Feron et al., 1981; Maltoni et al., 1981; Til et al., 1991). It induced hepatic angiosarcomas in two experiments, lung angiosarcomas in one experiment and hepatocellular adenomas and hepatocellular carcinomas in two experiments.

    3.3 Subcutaneous and intraperitoneal injection

    When vinyl chloride was tested in rats by subcutaneous injection and by intraperitoneal injection in single studies, no increase in tumour incidence was observed (Maltoni et al., 1981).

    3.4 Transplacental administration and perinatal exposure

    The transplacental carcinogenicity of vinyl chloride was evaluated in one study in the offspring of rats exposed by inhalation on days 12–18 of pregnancy. A low incidence of tumours was observed in prenatally exposed offspring at several sites including the kidney (nephroblastomas) and the Zymbal gland (carcinomas). However, no angiosarcomas or hepatomas developed in the offspring (Maltoni et al., 1981).

    Vinyl chloride was tested by perinatal inhalation exposure in two studies in rats. In one study, rats were exposed transplacentally, as neonates, and during adulthood. Treatment with vinyl chloride induced hepatic angiosarcomas and hepatocellular carcinomas. The rats also showed

    458

  • Table 3.1 Carcinogenicity studies in experimental animals exposed to vinyl chloride and chloroethylene oxide

    Species, strain (sex) Dosing regimen, Incidence of tumours Significance Comments Duration Animals/group at start Reference

    Mouse, NMRI (M, F) 26 or 52 wk Holmberg et al. (1976)

    Inhalation, 0, 50, 500 ppm [0, 130, 1 300 mg/ m3] 6 h/d, 5 d/wk, for 26 wk (0, 500 ppm) or 52 wk, (0, 50 ppm) 12/group/sex

    Extrahepatic angiosarcomas: (M) 0/24, 6/12*, 3/12*; (F) 0/24, 8/12*, 5/12*

    NR, *[p 

  • Table 3.1 (continued)

    Species, strain (sex) Dosing regimen, Incidence of tumours Significance Comments Duration Animals/group at start Reference

    Haemangiosarcomas (all sites): 1/71, 29/67, 30/47, 20/45 Lung carcinomas: 9/71, 18/65, 15/47, 11/45 Mammary gland carcinomas: 2/71, 33/67, 22/47, 22/45

    Study 2: 8 or 14 mo old mice Study 2, incidence for controls and Study 2, *p 

  • Table 3.1 (continued)

    Species, strain (sex) Dosing regimen, Incidence of tumours Significance Comments Duration Animals/group at start Reference

    Mouse CD-1 (M) up to 44–45 wk Suzuki (1983)

    Inhalation 0, 1, 10, 100, 300, 600 ppm [0, 2.6, 26, 260, 780, 1560 mg/m3] 6 h/d, 5 d/wk, 4 wk 30/group except 60/control group and 40/600 ppm-treated group

    Benign pulmonary tumours: NR *[p 

  • Table 3.1 (continued)

    Species, strain (sex) Dosing regimen, Incidence of tumours Significance Comments Duration Animals/group at start Reference

    IARC M

    ON

    OG

    RAPH

    S – 100F

    Rat, Sprague Dawley (M, F) 156 wk Maltoni et al. (1981)

    Inhalation 0, 50, 250, 500, 2500, 6000, 10000 ppm [130, 650, 1300, 6500, 15600, 26000 mg/m3] 4 h/d, 5 d/wk, 17 wk 30/group/sex, 190 controls

    Liver angiosarcomas: 0/190, 0/58, 0/59, 1/60, 1/60, 1/60, 0/58

    NR [p 

  • Table 3.1 (continued)

    Species, strain (sex) Dosing regimen, Incidence of tumours Significance Comments Duration Animals/group at start Reference

    Rat, Fischer (F) Lifetime Drew et al. (1983)

    Inhalation 0, 100 ppm [0, 260 mg/m3] 6 h/d, 5 d/wk, for 6, 12, 18 or 24 mo 55–112/group

    Tumour incidence after 24 mo of exposure: *p 

  • Table 3.1 (continued)

    Species, strain (sex) Dosing regimen, Incidence of tumours Significance Comments Duration Animals/group at start Reference

    Hepatocellular adenomas: (M) 0/99, 0/99, 0/99, 1/49; (F) 0/98, 1/100, 1/96, 9/49* Liver haemangiosarcomas: (M) 0/99, 0/99, 0/99, 1/49; (F) 0/98, 0/100, 0/96, 2/49

    Rat, Sprague-Dawley Subcutaneous injection Mammary gland tumours (malignant): 3/75, NR, [NS] 99.97% purity (M, F) Single injection of 0, 4.25 mg in 1 1/75. Nephroblastomas: 0/75, 1/75 No other tumour types 145 wk ml olive oil, observed Maltoni et al. (1981) 35 M/group, 40 F/group Data reported for both sexes

    combined Rat, Sprague-Dawley Intraperitoneal injection Extrahepatic angiosarcomas: 0/55, 0/55, 1/56, NR, [NS] 99.97% purity (M, F) 0 (olive oil, once); 4.25 mg/kg bw, 1/53, 0/56 No liver angiosarcomas 144 wk Maltoni et al. (1981)

    once, twice, three, or four times at 2 mo intervals 30/group/sex

    Mammary gland tumours (malignant): 0/55, 2/55, 3/56, 1/53, 1/56

    observed Data reported for both sexes combined

    Rat, Sprague-Dawley Transplacental In offspring - 99.97% purity (M, F) Pregnant females exposed by Extrahepatic angiomas: 1/32, 0/51 No angiosarcomas (all sites) 143 wk inhalation to 6000, 10000 ppm Nephroblastomas: 0/32, 3/51 or hepatomas were observed Maltoni et al. (1981) [15600, 26000 mg/m3] Zymbal gland carcinomas: 3/32, 5/51 in offspring. One Zymbal

    on D 12–18 of pregnancy Skin epitheliomas: 1/32, 0/51 gland tumour at high dose 30–54 dams/group Forestomach papillomas and achanthomas: (1/30) was the only tumour

    1/32, 1/51 observed in dams. Despite Mammary gland tumours (malignant): 2/32, the lack of controls, this study 1/51 provides some evidence of the

    transplacental carcinogenicity of vinyl chloride. Data reported for both sexes combined.

    IARC M

    ON

    OG

    RAPH

    S – 100F

    Rat, Wistar (M, F) 149 wk (M), 150 wk (F) Til et al. (1991)

    Oral (feed) PVC powder in food that resulted in dose of 0, 0.014, 0.13, 1.3 mg vinyl chloride/kg bw 4–6 h/d, 7 d/wk 50–100/group/sex

    Hepatocellular carcinomas: (M) 0/99, 0/99, 0/99, 3/49*; (F) 1/98, 0/100, 1/96, 3/49

    *p 

  • Table 3.1 (continued)

    Species, strain (sex) Dosing regimen, Incidence of tumours Significance Comments Duration Animals/group at start Reference

    Rat, Sprague-Dawley (M, F) 124 wk Maltoni et al. (1981)

    Perinatal Breeders and newborn offspring exposed together by inhalation to. 6000, 10000 ppm [15600, 26000 mg/m3] 4 h/d, 5 d/wk, 5 wk 42–44/group (offspring)

    In offspring Hepatic angiosarcomas: 17/42, 15/42 Liver angiomas: 1/42, 0/44 Extrahepatic angiosarcomas: 1/42, 0/44 Extrahepatic angiomas: 1/42, 3/44 Hepatomas: 20/42, 20/44 Zymbal gland carcinomas: 2/42, 1/44 Skin epitheliomas: 2/42, 1/44 Mammary gland tumours (malignant): 1/42, 0/44

    - 99.97% No concurrent controls No tumours observed in breeders at sites where tumours occurred in offspring Data reported for both sexes combined

    Rat, Sprague-Dawley (M, F) Lifetime Maltoni & Cotti (1988)

    Hamster, Golden (M) 109 wk Maltoni et al. (1981)

    Perinatal Breeders exposed by inhalation to 0, 2500 ppm [0, 6500 mg/m3], 4 h/d, 5 d/wk, 7 wk; dams became pregnant and delivered offspring. Dams exposed (7 h/d) for 69 additional wk with group I offspring; group II offspring exposed only for 8 additional wk 54–60 breeders (F)/group, 60–63 offspring/group/sex, 149–158 control offspring/group/ sex Inhalation 0, 50, 250, 500, 2500, 6000, 10 000 ppm [0, 130, 650, 1300, 6500, 15600, 26000 mg/m3] 4 h/d, 5 d/wk, 30 wk 30/group, 60 controls

    Hepatocarcinomas: Breeders (F): 0/60, 5/54 NR 99.97% purity Offspring group I: (M) 27/64; (F) 38/63; [p 

  • Table 3.1 (continued)

    Species, strain (sex) Dosing regimen, Incidence of tumours Significance Comments Duration Animals/group at start Reference

    Mammary gland carcinomas: 0/143, 28/87**, 31/52**, 47/102**, 2/52*, 0/50, 1/52, 31/52**, 6/44**, 0/42 Stomach adenomas: 5/138, 23/88**, 3/50*, 20/101**, 15/53**, 6/49*, 0/52, 3/50*, 10/44**, 3/41 Skin carcinomas: 0/133, 2/80, 9/48**, 3/90, 0/49, 0/46, 0/50, 2/80, 0/38, 0/30

    IARC M

    ON

    OG

    RAPH

    S – 100F

    Hamster, Syrian golden (F)

    Inhalation 200 ppm [520 mg/m3] 13/88**, 4/52**, 2/103, 3/53*, 0/50, 0/52, **p 

  • Vinyl chloride

    a high incidence of tumours that were probably of olfactory neuroepithelial origin, but were formerly reported as cerebral neuroblastomas in some studies. Similar results were observed in co-exposed dams (Maltoni & Cotti, 1988). In a second study, rats were exposed to vinyl chloride for five weeks, beginning at birth. Angiosarcomas of the liver and hepatomas occurred at a high incidence in the offspring, but not in the dams that were co-exposed with the offspring (Maltoni et al., 1981).

    3.5 Carcinogenicity of metabolites

    Chloroethylene oxide, a chemically reactive metabolite of vinyl chloride, was tested for carcinogenicity in a single study in mice by subcutaneous injection and in an initiation–promotion protocol by skin application. It caused a massive increase of fibrosarcomas at the site of injection and increased the incidence of squamous-cell papillomas and carcinomas of the skin at the site of application (Zajdela et al., 1980).

    4. Other Relevant Data

    4.1 Kinetics and metabolism – studies in humans

    Pulmonary absorption of vinyl chloride in humans appears to be rapid and the percentage absorbed is independent of the concentration inhaled. Adult male volunteers exposed for six hours to air containing 2.9–23.1 ppm [7.5–60 mg/m3] vinyl chloride, retained on average approximately 42% of the inhaled amount (Krajewski et al., 1980; cited in ATSDR, 2006). Pulmonary uptake is determined in part by the blood–air partition constant, which is 1.16 for vinyl chloride (Gargas et al., 1989). Even if no data in humans were available, by assuming an identical solubility of vinyl chloride in rodent

    and human tissues, the tissue–blood partition constants would be twofold greater in humans (Clewell et al., 2001), as a consequence of the twofold lower blood–air partition coefficient of vinyl chloride in humans compared with rats and mice.

    In the postmitochondrial fractions of liver homogenates of humans and rats, large interindividual variations were noted in the metabolism of vinyl chloride, while the average activity was comparable between rat and human samples (Sabadie et al., 1980). Vinyl chloride is primarily and rapidly metabolized in the liver (see Fig. 4.1), with a saturable mechanism (Reynolds et al., 1975; Ivanetich et al., 1977; Barbin & Bartsch, 1989; Lilly et al., 1998; Bolt, 2005). The first step is oxidation in the liver, predominantly mediated by the human cytochrome P450 (CYP) isoenzyme 2E1 (WHO, 1999). Since CYP2E1 is present in several tissues at low levels – compared with concentrations in the liver – extrahepatic metabolism of systemically available vinyl chloride does occur. Inhibitors of CYP, such as 3-bromophenyl-4(5)-imidazole or 6-nitro1,2,3-benzothiadiazole, reduce the metabolism of vinyl chloride in vivo (Bolt et al., 1976). The primary metabolites of vinyl chloride are the highly reactive chloroethylene oxide, which is formed in a dose-dependent process and has a half-life of 1.6  minutes in aqueous solution at neutral pH (Barbin et al., 1975; Dogliotti, 2006), and its rearrangement product chloroacetaldehyde (Bonse et al., 1975). Both can bind to proteins, DNA and RNA and form ethenoadducts; chloroethylene oxide is the most reactive with nucleotides (Guengerich et al., 1979).

    Conjugation of chloroethylene oxide and chloroacetaldehyde with glutathione (GSH) eventually leads to the major urinary metabolites N-acetyl-S-(2-hydroxyethyl)cysteine and thiodiglycolic acid (Plugge & Safe, 1977). The latter compound has been reported to be the major metabolite in the urine of exposed workers (Cheng et al., 2001) with concentrations in

    467

  • IARC MONOGRAPHS – 100F

    Fig. 4.1 Proposed metabolic pathways for vinyl chloride ClHC CH2 Vinyl chloride

    CYP2E1 mixed-function oxidase

    DNA adducts

    OProtein adducts H2C CH

    O Cl H2O HCl O Re-arrangement HO CH2 C Epoxide hydrolase Glycolaldehyde

    ClH2C C 2-Chloroethylene oxide

    H HGlutathione 2-Chloroacetaldehyde

    Glutathione

    Aldehyde O dehydrogenase G S CH2 C

    O S-Formylmethyl glutathione H

    ClH2C C

    OH 2-Chloroacetic acid (u)

    O Glutathione cys-S CH2 C

    HS-Formylmethyl cysteine (u)

    G S CH2 COOH S-Carboxymethyl glutathione

    cys-S CH2 CH2OH cys-S CH2 COOH S-(2-Hydroxyethyl)cysteine

    S-Carboxymethyl cysteine (u)

    NH3

    N-Ac-cys-S CH2 CH2OH(Transamination) N-Acetyl-S-(2-Hydroxyethyl)cysteine (u)

    CO2

    (Oxidative decarboxylation)

    Major pathway Minor pathway

    HOOC CH2 S CH2 COOH Thiodiglycolic acid (thiodiacetic acid) (u)

    From Barbin et al. (1975), Plugge & Safe (1977), Green & Hathway (1977), Guengerich & Watanabe (1979), Guengerich et al. (1979), Bolt et al.

    (1980), adapted from ATSDR (2006). CYP, cytochrome P450; (u), excreted in urine

    468

  • Vinyl chloride

    urine that correlated with environmental vinyl chloride concentrations of >  5 ppm (ATSDR, 2006). Chloroethylene oxide can also be detoxified to glycolaldehyde by microsomal epoxide hydrolase (mEH), while chloroacetaldehyde can be converted to chloroacetic acid by aldehyde dehydrogenase 2 (ALDH2) in the urine (Guengerich & Watanabe, 1979; ATSDR, 2006; IARC, 2008). Another route of elimination of vinyl chloride is exhalation of the unmetabolized compound, which occurs at low levels in humans (Müller et al., 1978; Krajewski et al., 1980; Pleil & Lindstrom, 1997). When volunteers were exposed for six hours to air containing 6.8–23.1 ppm [15–60 mg/m3] vinyl chloride, the mean concentration in exhaled air ranged from 0.21 to 1.11 ppm [0.54–2.84 mg/m3], representing 3.6 and 4.7%, respectively, of the inhaled amount of vinyl chloride (Krajewski et al., 1980).

    4.2 Kinetics and metabolism – studies in animals

    Experimental studies on vinyl chloride have been evaluated in previous IARC Monographs (IARC, 1979, 1987, 2008). Comprehensive data on the mechanism of vinyl chloride-induced carcinogenicity are available, encompassing toxicokinetics, metabolism, biomarkers, and genotoxicity. Many key events in the pathway of vinyl chloride-induced hepatocarcinogenesis have been established (Bolt, 2005; Dogliotti, 2006; IARC, 2008).

    The absorption, distribution, metabolism and elimination of vinyl chloride in rats and mice have been reviewed in IARC Monograph Volumes 19 and 97 (IARC, 1979, 2008) and elsewhere (WHO, 1999; ATSDR, 2006); the most relevant data are summarized below.

    In animals, pulmonary and gastrointestinal absorption of vinyl chloride occurs readily and rapidly, while dermal absorption is probably not significant. In monkeys exposed (whole body,

    except the head) to atmospheres containing 7000 and 800 ppm vinyl chloride for, respectively, 2.0 and 2.5  hours, only 0.023–0.031% of the total available amount of vinyl chloride was absorbed via the dermal route (Hefner et al., 1975a), whereas intestinal absorption and uptake in blood was virtually complete in 10 minutes in rats after single oral doses (44–92 mg/kg bw) in aqueous solution (Withey, 1976). Pulmonary absorption in rats amounted to about 40% of inhaled [14C]-labelled vinyl chloride for initial exposure concentrations below 260 mg/m3 [100 ppm] (Bolt et al., 1976).

    The tissue–blood partition constants determine the distribution volume of vinyl chloride, and range from 0.4 (muscle) to 10 (fat) in male rats (Barton et al., 1995). The fat–air partition constant for vinyl chloride, reported by several authors, tends to be higher in females than in males (WHO, 1999). Following inhalation, vinyl chloride is distributed in several tissues such as brain, liver, spleen, kidney, adipose tissue and muscle, with the highest levels found in liver and kidney (Bolt et al., 1976). It has also been detected in fetal blood and amniotic fluid of rats after a 2.5-hour exposure to ~2000–13000 ppm [5200–33800 mg/m3]), which indicates its capability to cross the placental barrier (Ungváry et al., 1978).

    CYP2E1 appears to account for all metabolic activity in rat liver microsomes, with a maximum velocity (Vmax) of 4674 pmol/mg protein/min and a Michaelis-Menten constant (Km) of 7.42 μmol/L (El Ghissassi et al., 1998). Chloroacetic acid was metabolized in rats to two major urinary metabolites, viz. S-(carboxymethyl) cysteine and thiodiacetic acid (Yllner, 1971). S-(carboxymethyl)cysteine, S-(2-chloroethyl) cysteine and N-acetyl-S-vinylcysteine are metabolites of vinyl chloride in rats after oral administration (Watanabe et al., 1976a; Green & Hathway, 1975 and 1977) and N-acetyl-S-(2hydroxyethyl)cysteine is a metabolite after inhalation (Watanabe et al., 1976b). Thiodiglycolic

    469

    http:0.54�2.84

  • IARC MONOGRAPHS – 100F

    acid was obtained as a common metabolite in rats dosed separately with either chloroacetaldehyde, chloroacetic acid or S-(carboxymethyl) cysteine. Therefore, the identification of the same S-containing metabolite from vinyl chloride-treated animals lends support to the hypothesis that chloroethylene oxide or chloroacetaldehyde are formed and react with GSH (Green & Hathway, 1977). Following oral administration of [14C]-labelled vinyl chloride to rats, [14C]-carbon dioxide (Green & Hathway, 1975; Watanabe et al., 1976a), [14C]-labelled urea and glutamic acid were identified as minor metabolites (Green & Hathway, 1975).

    Saturation of the metabolism of vinyl chloride (Gehring et al., 1978; Filser & Bolt, 1979) appears to occur at inhalation concentrations above 200 ppm [520 mg/m3] in rhesus monkeys (Buchter et al., 1980) and above 250 ppm [650 mg/m3] in rats (Bolt et al., 1977; Filser & Bolt, 1979). The plateau of incidence of ASL in carcinogenicity bioassays is also observed in rats at exposures above 250 ppm (reviewed in Bolt, 2005).

    ATSDR (2006) summarized the kinetic constants obtained in vivo in male Sprague-Dawley rats (Vmax, 58 μmol/h/kg; Km, 1 μM) and rhesus monkeys (Vmax, 50 μmol/h/kg) (based on Buchter et al., 1980; Barton et al., 1995). The latter value (50 μmol/h/kg) was suggested to be a closer approximation to human metabolism than the value of 110 μmol/h/kg estimated for rats by Filser & Bolt (1979) (ATSDR, 2006).

    Watanabe et al. (1978a) reported that the elimination rate of vinyl chloride was not altered during repeated exposures via inhalation (five days per week during seven weeks) compared with a single inhalation exposure (~13 000 mg/m3 [5000 ppm]).

    Urinary excretion of polar metabolites of vinyl chloride is the predominant route of elimination at low concentrations, and only very small amounts are expired in the air unchanged (Hefner et al., 1975b). Once metabolic saturation is attained, vinyl chloride is eliminated via other

    routes, mainly exhalation of the parent chemical. Following exposure of male rats to 26 mg/m3 [10 ppm] [14C]-labelled vinyl chloride by inhalation during six hours, urinary [14C] radioactivity and expired vinyl chloride (measured as [14C]-labelled carbon dioxide) were recovered in amounts of 68% and 2%, respectively; after exposure to a 100-fold higher concentration, the proportion of radioactivity in the urine decreased to 56% and the amount expired increased to 12% (Watanabe et al., 1976b). Moreover, the same authors showed that after single oral doses of 0.05, 1 or 100 mg/kg bw [14C]-labelled vinyl chloride, urinary excretion of radioactivity was 68, 59 and 11%, respectively; expired [14C]-labelled carbon dioxide accounted for 9, 13 and 3%, respectively; pulmonary elimination of [14C]-labelled vinyl chloride represented only 1–3% of the lower dose and 67% of the higher dose (Watanabe et al., 1976a). The route of elimination may depend upon the route of administration, since urinary excretion is favoured after oral or intra-peritoneal administration, which indicates a first-pass effect due to metabolism in the liver (Clewell et al., 2001).

    4.3 Reaction with cellular macromolecules

    Vinyl chloride is a genotoxic carcinogen in animals and humans (Block, 1974; Creech & Johnson, 1974; Lee & Harry, 1974; Maltoni et al., 1974, 1981). It is mutagenic, usually in the presence of metabolic activation, in various assays with bacteria, yeast or mammalian cells; it is also clastogenic in vivo and in vitro. Vinyl chloride induces unscheduled DNA synthesis, increases the frequency of sister chromatid exchange in rat and human cells, and increases the frequency of chromosomal aberrations and micronucleus formation in mice, rats, and hamsters in vivo (IARC, 2008).

    Osterman-Golkar et al. (1976) reported the alkylation of haemoglobin – at cysteine and

    470

  • Vinyl chloride

    histidine – and small amounts of alkylated histidine in proteins from the testis of mice exposed to [14C]-labelled vinyl chloride. Binding of nonvolatile metabolites of [14C]-labelled vinyl chloride to liver macromolecules has been observed, both in vitro and in rats exposed by inhalation (Kappus et al., 1976; Watanabe et al., 1978a, b; Guengerich & Watanabe, 1979; Guengerich et al., 1979; Bolt et al., 1980; Guengerich et al., 1981; Barton et al., 1995). A decrease in non-protein sulfhydryl concentration was seen in rats after exposure to high concentrations of vinyl chloride (Jedrychowski et al. (1984). Kappus et al. (1975) and Laib & Bolt (1977) reported binding of vinyl chloride to RNA in an in-vitro incubation with rat-liver microsomes, and to liver RNA of rats exposed in vivo. Watanabe et al. (1978b) reported macromolecular binding proportional to the amount of vinyl chloride metabolized, but not proportional to the exposure concentration.

    Chloroethylene oxide and chloroacetaldehyde can form etheno adducts with nucleic acid bases in vitro (Guengerich, 1992). Chloroethylene oxide yields the N7-(2oxoethyl)guanine adduct (7-OEG), four etheno adducts – 1,N6-ethenoadenine (εA), 3,N4-ethenocytosine (εC), N2,3-ethenoguanine (N2,3-εG) and 1,N2-ethenoguanine (1,N2-εG) (Ciroussel et al., 1990; Guengerich, 1992), and 5,6,7,9-tetrahydro-7-hydroxy-9-oxoimidazo[1,2a]purine (HO-ethanoG) (Müller et al., 1996). In rats, the DNA adducts εA and εC have been found in various organs after exposure to vinyl chloride by inhalation. 7-OEG was the major DNA adduct formed in vivo and was found in greater amounts in young animals (Swenberg et al., 2000). However, 7-OEG has a short half-life of about 62 hours, while the etheno adducts are more persistent. For example, N2,3-εG (which is 10–100-fold more abundant than other etheno adducts in exposed animals) has a half-life of about 30 days (Fedtke et al., 1990). After exposure of rats to 500 ppm [1300 mg/m3] vinyl chloride for eight weeks, the εA level was significantly

    increased above background in liver, lung, lymphocytes and testis, while the amount of εC was increased in liver, kidney, lymphocytes and spleen, but not in brain (Guichard et al., 1996; Barbin, 1999). When adult rats were exposed to 1100 ppm [2860 mg/m3] vinyl chloride for one or four weeks, there was a significant increase in the level of N2,3-εG in hepatocytes and nonparenchymal cells, but not in the brain, with a linear increase at exposure concentrations from 0 to 100 ppm [260 mg/m3] and a plateau at 100–1100 ppm [260–2860 mg/m3]. In weanling animals there was a small, statistically significant increase in N2,3-εG in the brain after five days of exposure and the amount of N2,3-εG in hepatocytes was significantly greater than that measured in non-parenchymal cells after exposures to 10 and 100 ppm [26 and 260 mg/m3] vinyl chloride (Morinello et al., 2002a). These differential responses between weanlings and adults may contribute to the particular susceptibility of young rats to vinyl chloride-induced neuroblastomas and HCC (Maltoni & Cotti, 1988). There was no significant difference in N2,3-εG-adduct levels, nor in the rate of repair between hepatocytes and non-parenchymal cells (Morinello et al., 2002b), which confirms the earlier observation of Yang et al. (2000). Data on the occurrence and persistence of vinyl chloride-DNA adducts in humans are still lacking. Nair et al. (1995) used immunoaffinity purification of the etheno adducts and subsequent [32P]-postlabelling, and reported values of 14.1 εA and 8.1 εC per 109 parent bases in non-neoplastic liver tissue of a vinyl chloride-exposed patient with HCC. These adducts may also result from lipid peroxidation (El Ghissassi et al., 1995) and their level can be quite high in patients with unknown exposure (up to 0.5–40 εA and εC per 109 parent bases in liver-DNA samples) (Bartsch & Nair, 2000a, b).

    Barbin et al. (1985) reported that the 7-OEG– DNA adduct lacks miscoding or promutagenic properties. In contrast, etheno adducts and related exocyclic DNA adducts (e.g. εΑ, εC,

    471

  • IARC MONOGRAPHS – 100F

    N2,3−εG, and HO-ethanoG) may be involved in base-pair substitution and other specific mutations in cancer-related genes (i.e. RAS oncogenes, Tp53 tumour-suppressor gene) (WHO, 1999). The DNA lesions εA, εC and N2,3-εG have demonstrated miscoding potential in vitro and in vivo (Singer et al., 1987; Cheng et al., 1991; Mroczkowska & Kuśmierek, 1991; Singer et al., 1991; Basu et al., 1993). The adduct εA causes A→G transitions and A→T transversions, εC causes C→A transversions and C→T transitions and εG causes G→A transitions (Bolt, 2005). The same mutation types are observed in Tp53 and RAS genes in vinyl chloride-induced tumours. Mutations in Ki-RAS are associated with vinyl chloride-induced angiosarcomas in humans but not in rats, and to a lesser extent with vinyl chloride-induced HCC (CAA61CTA Ha-Ras mutation) in rats (IARC, 2008). In half of the cases, these mutations led to the incorporation of aspartate instead of glycine. Tp53 mutations associated with exposure to vinyl chloride (frequently A→T transversions) are found in approximately half of the angiosarcomas in both humans and rats. The presence of mutated p21ras and p53 proteins in the blood of a high proportion of workers exposed to vinyl chloride and the positive correlation between the occurrence of the mutated proteins and cumulative exposure to vinyl chloride, suggest that the mutation is an early event (IARC, 2008).

    Various assays have been designed to explore the mutagenic properties of DNA adducts introduced into oligonucleotides or into site-specific vectors. Vector plasmids have also been treated with 2-chloroethyleneoxide or 2-chloroacetaldehyde and propagated in E. coli or mammalian cells. The mechanism by which adducts cause mutations still remains unclear, as misincorporation events depend on the individual mechanisms of DNA polymerases (Choi et al., 2006). HO-ethanoG and 1,N2-εG block the replication process with many different polymerases, thereby

    causing base misincorporation (Langouët et al., 1997, 1998; Guengerich et al., 1999).

    The induction of extrahepatic tumours (e.g. in the brain or lung) by vinyl chloride has been established experimentally, but the mechanism is not well elucidated (Bolt, 2005). Overall, data suggest that etheno adducts are probably involved in the initiation of hepatocarcinogenesis, but the effects of the observed tissue- and cell-specificity and the variability in various biomarkers such as mutant p53 and anti-p53 antibodies are not completely clear (Trivers et al., 1995; Brandt-Rauf et al., 1996). One source for this variability may be explained by differences in polymorphisms in genes (i.e. CYp2E1, GSTT1, GSTM1, ALDH2) that encode metabolising enzymes or DNA-repair proteins (i.e. the XRCC1 gene) (Li et al., 2003a).

    4.4 Synthesis

    Numerous studies on the toxicokinetics, metabolism, genotoxicity, and molecular biology of vinyl chloride provide strong evidence that the carcinogenicity of this chamical involves a genotoxic mechanism of action, mediated by reactive metabolites. The extensive information on the mechanism underlying vinyl chloride-induced carcinogenicity has established many key events in the pathway of vinyl chloride-induced liver carcinogenesis. These key events include metabolic activation to reactive metabolites, binding of the metabolites to DNA, promutagenic action of these adducts leading to G→A and A→T transitions, and the effects of such mutations on the functioning of proto-oncogenes and tumoursuppressor genes at the gene and protein levels, with tumourigenesis as the final outcome. Many of these key events identified in experimental animals have also been demonstrated in humans.

    472

  • Vinyl chloride

    5. Evaluation

    There is sufficient evidence in humans for the carcinogenicity of vinyl chloride. Vinyl chloride causes angiosarcoma of the liver, and hepatocellular carcinoma.

    There is sufficient evidence in experimental animals for the carcinogenicity of vinyl chloride.

    There is sufficient evidence in experimental animals for the carcinogenicity of chloroethylene oxide.

    There is strong evidence that the carcinogenicity of vinyl chloride operates by a genotoxic mechanism that involves metabolic activation to reactive metabolites, binding of the metabolites to DNA, promutagenic action of these adducts leading to mutations in proto-oncogenes and tumour-suppressor genes. Many of these key events identified in experimental animals have also been demonstrated in humans.

    Vinyl chloride is carcinogenic to humans (Group 1).

    References

    ATSDR (2006). Toxicological profile on Vinyl Chloride. Atlanta, GA: Centers for Disease Control and Prevention. Available at http://www.atsdr.cdc.gov/ toxprofiles/phs20.html

    Barbin A (1999). Role of etheno DNA adducts in carcinogenesis induced by vinyl chloride in rats. IARC Sci publ, 150: 303–313. PMID:10626230

    Barbin A & Bartsch H (1989). Nucleophilic selectivityas a determinant of carcinogenic potency (TD50) inrodents: a comparison of mono- and bi-functionalalkylating agents and vinyl chloride metabolites. Mutat Res, 215: 95–106. PMID:2811916

    Barbin A, Brésil H, Croisy A et  al. (1975). Livermicrosome-mediated formation of alkylating agentsfrom vinyl bromide and vinyl chloride. Biochem Biophys Res Commun, 67: 596–603. doi:10.1016/0006291X(75)90854-2 PMID:1201042

    Barbin A, Laib RJ, Bartsch H (1985). Lack of miscodingproperties of 7-(2-oxoethyl)guanine, the major vinylchloride-DNA adduct. Cancer Res, 45: 2440–2444. PMID:3986785

    Barton HA, Creech JR, Godin CS et  al. (1995).Chloroethylene mixtures: pharmacokinetic modelingand in vitro metabolism of vinyl chloride, trichloroethylene, and trans-1,2-dichloroethylene in rat. Toxicol Appl pharmacol, 130: 237–247. doi:10.1006/taap.1995.1029 PMID:7871537

    Bartsch H & Nair J (2000a). New DNA-based biomarkersfor oxidative stress and cancer chemoprevention studiesEuropean Journal of Cancer (Oxford, England), 36: 1229–1234.

    Bartsch H & Nair J (2000b). Ultrasensitive and specificdetection methods for exocylic DNA adducts: markersfor lipid peroxidation and oxidative stress. Toxicology, 153: 105–114. PMID:11090950

    Basu AK, Wood ML, Niedernhofer LJ et  al. (1993).Mutagenic and genotoxic effects of three vinyl chloride-induced DNA lesions: 1,N6-ethenoadenine, 3,N4-ethenocytosine, and 4-amino-5-(imidazol-2-yl)imidazole. Biochemistry, 32: 12793–12801. doi:10.1021/bi00210a031 PMID:8251500

    Block JB (1974). Angiosarcoma of the liver following vinylchloride exposure. JAMA, 229: 53–54. doi:10.1001/jama.229.1.53 PMID:4406736

    Boffetta P, Matisane L, Mundt KA, Dell LD (2003). Meta-analysis of studies of occupational exposure to vinylchloride in relation to cancer mortality. Scand J Work Environ Health, 29: 220–229. PMID:12828392

    Bolt HM (2005). Vinyl chloride-a classical industrialtoxicant of new interest. Crit Rev Toxicol, 35: 307–323. doi:10.1080/10408440490915975 PMID:15989139

    Bolt HM, Filser JG, Laib RJ, Ottenwälder H (1980). Bindingkinetics of vinyl chloride and vinyl bromide at very lowdoses. Arch Toxicol Suppl, 3: 129–142. PMID:6930940

    Bolt HM, Kappus H, Buchter A, Bolt W (1976). Dispositionof (1,2–14C) vinyl chloride in the rat. Arch Toxicol, 35: 153–162. doi:10.1007/BF00293562 PMID:989287

    Bolt HM, Laib RJ, Kappus H, Buchter A (1977).Pharmacokinetics of vinyl chloride in the rat. Toxicology, 7: 179–188. doi:10.1016/0300-483X(77)90063-4 PMID:857343

    Bonse G, Urban T, Reichert D, Henschler D (1975).Chemical reactivity, metabolic oxirane formation and biological reactivity of chlorinated ethylenes inthe isolated perfused rat liver preparation. Biochem pharmacol, 24: 1829–1834. doi:10.1016/00062952(75)90468-2 PMID:1233985

    Boraiko C & Batt JTin Stabilizers Association (2005).Evaluation of employee exposure to organic tin compounds used as stabilizers at PVC processingfacilities. J Occup Environ Hyg, 2: 73–76, quiz D6–D7.doi:10.1080/15459620590906810 PMID:15764527

    Brandt-Rauf PW, Chen JM, Marion MJ et  al. (1996).Conformational effects in the p53 protein of mutationsinduced during chemical carcinogenesis: moleculardynamic and immunologic analyses. J protein Chem, 15: 367–375. doi:10.1007/BF01886863 PMID:8819013

    473

    http://www.ncbi.nlm.nih.gov/pubmed/10626230http://www.ncbi.nlm.nih.gov/pubmed/2811916http://dx.doi.org/10.1016/0006-291X(75)90854-2http://dx.doi.org/10.1016/0006-291X(75)90854-2http://www.ncbi.nlm.nih.gov/pubmed/1201042http://www.ncbi.nlm.nih.gov/pubmed/3986785http://dx.doi.org/10.1006/taap.1995.1029http://dx.doi.org/10.1006/taap.1995.1029http://www.ncbi.nlm.nih.gov/pubmed/7871537http://www.ncbi.nlm.nih.gov/pubmed/11090950http://dx.doi.org/10.1021/bi00210a031http://dx.doi.org/10.1021/bi00210a031http://www.ncbi.nlm.nih.gov/pubmed/8251500http://dx.doi.org/10.1001/jama.229.1.53http://dx.doi.org/10.1001/jama.229.1.53http://www.ncbi.nlm.nih.gov/pubmed/4406736http://www.ncbi.nlm.nih.gov/pubmed/12828392http://dx.doi.org/10.1080/10408440490915975http://www.ncbi.nlm.nih.gov/pubmed/15989139http://www.ncbi.nlm.nih.gov/pubmed/6930940http://dx.doi.org/10.1007/BF00293562http://www.ncbi.nlm.nih.gov/pubmed/989287http://dx.doi.org/10.1016/0300-483X(77)90063-4http://www.ncbi.nlm.nih.gov/pubmed/857343http://dx.doi.org/10.1016/0006-2952(75)90468-2http://dx.doi.org/10.1016/0006-2952(75)90468-2http://www.ncbi.nlm.nih.gov/pubmed/1233985http://dx.doi.org/10.1080/15459620590906810http://www.ncbi.nlm.nih.gov/pubmed/15764527http://dx.doi.org/10.1007/BF01886863http://www.ncbi.nlm.nih.gov/pubmed/8819013http:http://www.atsdr.cdc.gov

  • IARC MONOGRAPHS – 100F

    Buchter A, Filser JG, Peter H, Bolt HM (1980).Pharmacokinetics of vinyl chloride in the Rhesusmonkey. Toxicol Lett, 6: 33–36. doi:10.1016/03784274(80)90099-5 PMID:7423542

    CAREX (1999). Carex: industry specific estimates – Summary. Available at http://www.ttl.fi/en/chemical_safety/carex/Documents/5_exposures_by_agent_and_industry.pdf

    Casula D, Cherchi P, Spiga G, Spinazzola A (1977).Environmental dust in a plant for the production ofpolyvinyl chloride Ann Ist Super Sanita, 13: 189–198. PMID:603117

    CDC (1997). Epidemiologic notes and reports.Angiosarcoma of the liver among polyvinyl chlorideworkers–Kentucky. MMWR Morb Mortal Wkly Rep, 46: 97–101. PMID:9045039

    Cheng KC, Preston BD, Cahill DS et al. (1991). The vinylchloride DNA derivative N2,3-ethenoguanine producesG — -A transitions in Escherichia coli. proc Natl Acad Sci U S A, 88: 9974–9978. doi:10.1073/pnas.88.22.9974PMID:1946466

    Cheng TJ, Huang YF, Ma YC (2001). Urinary thiodiglycolic acid levels for vinyl chloride monomer-exposedpolyvinyl chloride workers. J Occup Environ Med, 43: 934–938. doi:10.1097/00043764-200111000-00002PMID:11725332

    Choi JY, Zang H, Angel KC et  al. (2006). Translesion synthesis across 1,N2-ethenoguanine by human DNA polymerases. Chem Res Toxicol, 19: 879–886. doi:10.1021/tx060051v PMID:16780368

    Ciroussel F, Barbin A, Eberle G, Bartsch H (1990). Investigations on the relationship between DNA ethenobase adduct levels in several organs of vinylchloride-exposed rats and cancer susceptibility. Biochem pharmacol, 39: 1109–1113. doi:10.1016/00062952(90)90291-R PMID:2322297

    Clewell HJ, Gentry PR, Gearhart JM et  al. (2001).Comparison of cancer risk estimates for vinyl chlorideusing animal and human data with a PBPK model.Sci Total Environ, 274: 37–66. doi:10.1016/S00489697(01)00730-6 PMID:11453305

    Cooper WC (1981). Epidemiologic study of vinyl chlorideworkers: mortality through December 31, 1972. Environ Health perspect, 41: 101–106. doi:10.2307/3429301 PMID:7199425

    Creech JL Jr & Johnson MN (1974). Angiosarcoma of liverin the manufacture of polyvinyl chloride. J Occup Med, 16: 150–151. PMID:4856325

    Dimmick WF (1981). EPA programs of vinyl chloridemonitoring in ambient air. Environ Health perspect, 41: 203–206. doi:10.2307/3429316 PMID:6895871

    Dogliotti E (2006). Molecular mechanisms of carcinogenesis by vinyl chloride. Ann Ist Super Sanita, 42: 163–169. PMID:17033136

    Drew RT, Boorman GA, Haseman JK et  al. (1983). Theeffect of age and exposure duration on cancer induction

    by a known carcinogen in rats, mice, and hamsters.Toxicol Appl pharmacol, 68: 120–130. doi:10.1016/0041008X(83)90361-7 PMID:6682580

    El Ghissassi F, Barbin A, Bartsch H (1998). Metabolic activation of vinyl chloride by rat liver microsomes:low-dose kinetics and involvement of cytochrome P4502E1. Biochem pharmacol, 55: 1445–1452. doi:10.1016/S0006-2952(97)00645-X PMID:10076537

    El Ghissassi F, Boivin S, Lefrançois L et  al. (1995).Glutathione transferase Mu1–1 (GSTM1) getotype inindividuals exposed to vinyl chloride: Relation to ‘vinylchloride disease’ and liver angiosarcoma. Clinical Chemistry, 41: 1922–1924.

    European Commission (2003). Integrated pollutionprevention and Control (IppC). Reference Document onBest Available Techniques in the Large Volume OrganicChemical Industry. Luxembourg.

    Fedtke N, Boucheron JA, Walker VE, Swenberg JA (1990).Vinyl chloride-induced DNA adducts. II: Formation and persistence of 7-(2-oxoethyl)guanine and N2,3ethenoguanine in rat tissue DNA. Carcinogenesis, 11: 1287–1292. doi:10.1093/carcin/11.8.1287 PMID:2387014

    Feron VJ, Hendriksen CF, Speek AJ et al. (1981). Lifespanoral toxicity study of vinyl chloride in rats. Food Cosmet Toxicol, 19: 317–333. doi:10.1016/0015-6264(81)90391-6PMID:7196371

    Feron VJ & Kroes R (1979). One-year time-sequenceinhalation toxicity study of vinyl chloride in rats.II. Morphological changes in the respiratory tract,ceruminous glands, brain, kidneys, heart and spleen.Toxicology, 13: 131–141. PMID:516073

    Feron VJ, Kruysse A, Til HP (1979). One-year time sequence inhalation toxicity study of vinyl chloridein rats. I. Growth, mortality, haematology, clinicalchemistry and organ weights. Toxicology, 13: 25–28. PMID:516069

    Filser JG & Bolt HM (1979). Pharmacokinetics of halogenated ethylenes in rats. Arch Toxicol, 42: 123–136. doi:10.1007/BF00316492 PMID:485853

    Fleig I & Thiess AM (1974). Chromosome tests in vinylchloride exposed workers] (Ger.). Arbeitsmedizin, Sozialmedizin, praventivmedizin, 12: 280–283.

    Gargas ML, Burgess RJ, Voisard DE et al. (1989). Partition coefficients of low-molecular-weight volatile chemicalsin various liquids and tissues. Toxicol Appl pharmacol, 98: 87–99. doi:10.1016/0041-008X(89)90137-3 PMID:2929023

    Gehring PJ, Watanabe PG, Park CN (1978). Resolutionof dose-response toxicity data for chemicals requiringmetabolic activation: example–vinyl chloride. Toxicol Appl pharmacol, 44: 581–591. doi:10.1016/0041008X(78)90266-1 PMID:567390

    German Environmental Office (1978). Air contaminationwith vinyl chloride (VC) from PVC products] (Ger.).Umweltbundes Ber, 5: 1–23.

    474

    http://dx.doi.org/10.1016/0378-4274(80)90099-5http://dx.doi.org/10.1016/0378-4274(80)90099-5http://www.ncbi.nlm.nih.gov/pubmed/7423542http://www.ncbi.nlm.nih.gov/pubmed/603117http://www.ncbi.nlm.nih.gov/pubmed/9045039http://dx.doi.org/10.1073/pnas.88.22.9974http://www.ncbi.nlm.nih.gov/pubmed/1946466http://dx.doi.org/10.1097/00043764-200111000-00002http://www.ncbi.nlm.nih.gov/pubmed/11725332http://dx.doi.org/10.1021/tx060051vhttp://www.ncbi.nlm.nih.gov/pubmed/16780368http://dx.doi.org/10.1016/0006-2952(90)90291-Rhttp://dx.doi.org/10.1016/0006-2952(90)90291-Rhttp://www.ncbi.nlm.nih.gov/pubmed/2322297http://dx.doi.org/10.1016/S0048-9697(01)00730-6http://dx.doi.org/10.1016/S0048-9697(01)00730-6http://www.ncbi.nlm.nih.gov/pubmed/11453305http://dx.doi.org/10.2307/3429301http://www.ncbi.nlm.nih.gov/pubmed/7199425http://www.ncbi.nlm.nih.gov/pubmed/4856325http://dx.doi.org/10.2307/3429316http://www.ncbi.nlm.nih.gov/pubmed/6895871http://www.ncbi.nlm.nih.gov/pubmed/17033136http://dx.doi.org/10.1016/0041-008X(83)90361-7http://dx.doi.org/10.1016/0041-008X(83)90361-7http://www.ncbi.nlm.nih.gov/pubmed/6682580http://dx.doi.org/10.1016/S0006-2952(97)00645-Xhttp://dx.doi.org/10.1016/S0006-2952(97)00645-Xhttp://www.ncbi.nlm.nih.gov/pubmed/10076537http://dx.doi.org/10.1093/carcin/11.8.1287http://www.ncbi.nlm.nih.gov/pubmed/2387014http://dx.doi.org/10.1016/0015-6264(81)90391-6http://www.ncbi.nlm.nih.gov/pubmed/7196371http://www.ncbi.nlm.nih.gov/pubmed/516073http://www.ncbi.nlm.nih.gov/pubmed/516069http://dx.doi.org/10.1007/BF00316492http://www.ncbi.nlm.nih.gov/pubmed/485853http://dx.doi.org/10.1016/0041-008X(89)90137-3http://www.ncbi.nlm.nih.gov/pubmed/2929023http://dx.doi.org/10.1016/0041-008X(78)90266-1http://dx.doi.org/10.1016/0041-008X(78)90266-1http://www.ncbi.nlm.nih.gov/pubmed/567390http://www.ttl.fi/en

  • Vinyl chloride

    Green T & Hathway DE (1975). The biological fate inrats of vinyl chloride in relation to its oncogenicity.Chem Biol Interact, 11: 545–562. doi:10.1016/00092797(75)90030-7 PMID:1201617

    Green T & Hathway DE (1977). The chemistry and biogenesis of the S-containing metabolites of vinyl chloride in rats. Chem Biol Interact, 17: 137–150. doi:10.1016/00092797(77)90080-1 PMID:328181

    Groth DH, Coate WB, Ulland BM, Hornung RW (1981).Effects of aging on the induction of angiosarcoma.Environ Health perspect, 41: 53–57. doi:10.2307/3429293 PMID:7199429

    Guengerich FP (1992). Roles of the vinyl chloride oxidation products 1-chlorooxirane and 2-chloroacetaldehyde in the in vitro formation of etheno adducts ofnucleic acid bases [corrected] Chem Res Toxicol, 5: 2–5. doi:10.1021/tx00025a001 PMID:1581532

    Guengerich FP, Crawford WM Jr, Watanabe PG (1979).Activation of vinyl chloride to covalently bound metabolites: roles of 2-chloroethylene oxide and 2-chloroacetaldehyde. Biochemistry, 18: 5177–5182. doi:10.1021/bi00590a023 PMID:497175

    Guengerich FP, Mason PS, Stott WT et al. (1981). Rolesof 2-haloethylene oxides and 2-haloacetaldehydesderived from vinyl bromide and vinyl chloride in irreversible binding to protein and DNA. Cancer Res, 41: 4391–4398. PMID:7030476

    Guengerich FP & Watanabe PG (1979). Metabolismof [14C]- and [36C]-labeled vinyl chloride in vivoand in vitro. Biochem pharmacol, 28: 589–596. doi:10.1016/0006-2952(79)90140-0 PMID:444246

    Guichard Y, el Ghissassi F, Nair J et al. (1996). Formation and accumulation of DNA ethenobases in adult Sprague-Dawley rats exposed to vinyl chloride.Carcinogenesis, 17: 1553–1559. PMID:8761409

    Guengerich FP, Langouët S, Mican AN et  al. (1999).Formation of etheno adducts and their effects on DNA polymerases. IARC Sci publ, 150: 137–145. PMID:10626215

    Hefner RE Jr, Watanabe PG, Gehring PJ (1975a). Percutaneous absorption of vinyl chloride. Toxicol Appl pharmacol, 34: 529–532. doi:10.1016/0041008X(75)90149-0 PMID:813333

    Hefner RE Jr, Watanabe PG, Gehring PJ (1975b).Preliminary studies on the fate of inhaled vinyl chloride monomer (VCM) in rats. Environ Health perspect, 11: 85–95. doi:10.2307/3428329 PMID:1175571

    Hoffmann D, Patrianakos C, Brunnemann KD, Gori GB (1976). Chromatographic determination of vinylchloride in tobacco smoke. Anal Chem, 48: 47–50. doi:10.1021/ac60365a063 PMID:1244767

    Holm L, Westlin A, Holmberg B (1982). Technical control measures in the prevention of occupational cancer. Anexample from the pVC industry. In: proc. Int. Symp. prevent. Occup. Cancer. Geneva: International Labour Office, pp. 538–546.

    Holmberg B, Kronevi T, Winell M (1976). The pathologyof vinyl chloride exposed mice. Acta Vet Scand, 17: 328–342. PMID:988738

    Hong CB, Winston JM, Thornburg LP et  al. (1981).Follow-up study on the carcinogenicity of vinylchloride and vinylidene chloride in rats and mice: tumour incidence and mortality subsequent to exposure. J Toxicol Environ Health, 7: 909–924. doi:10.1080/15287398109530034 PMID:7265317

    Huang M (1996). Epidemiological investigation on occupational malignant tumor in workers exposed tovinyl chloride. In: Ministry of public Health of China.National epidemiological study on eight occupationalcancers (1982–1984) [in Chinese]. Beijing: Ministry ofPublic Health of China, pp. 86–98.

    IARC (1974). Some anti-thyroid and related substances,nitrofurans and industrial chemical. IARC Monogr Eval Carcinog Risk Chem Man, 7: 1–326.

    IARC (1979). Some monomers, plastics and synthetic elastomers, and acrolein. IARC Monogr Eval Carcinog Risk Chem Hum, 19: 1–513. PMID:285915

    IARC (1987). Overall evaluations of carcinogenicity: anupdating of IARC Monographs volumes 1 to 42. IARC Monogr Eval Carcinog Risks Hum Suppl, 7: 1–440. PMID:3482203

    IARC (2004). Tobacco smoke and involuntary smoking.IARC Monogr Eval Carcinog Risks Hum, 83: 1–1438. PMID:15285078

    IARC (2008). 1,3-Butadiene, ethylene oxide and vinylhalides (vinyl fluoride, vinyl chloride and vinylbromide). IARC Monogr Eval Carcinog Risks Hum, 97: 1–510. PMID: 20232717.

    Infante PF, Petty SE, Groth DH et al. (2009). Vinyl chloride propellant in hair spray and angiosarcoma of theliver among hairdressers and barbers: case reports. Int J Occup Environ Health, 15: 36–42. PMID:19267125

    Ivanetich KM, Aronson I, Katz ID (1977). The interactionof vinyl chloride with rat hepatic microsomal cytochrome P-450 in vitro. Biochem Biophys Res Commun, 74: 1411–1418. doi:10.1016/0006-291X(77)90599-X PMID:14640

    Jedrychowski RA, Sokal JA, Chmielnicka J (1984).Influence of exposure mode on vinyl chloride action.Arch Toxicol, 55: 195–198. PMID:6497652

    Kappus H, Bold HM, Buchter A, Bolt W (1975). Rat livermicrosomes catalyse covalent binding of 14C-vinylchloride to macromolecules. Nature, 257: 134–135. PMID:240126

    Kappus H, Bolt HM, Buchter A, Bolt W (1976). Livermicrosomal uptake of (14C)vinyl chloride and transformation to protein alkylating metabolites in vitro.Toxicol Appl pharmacol, 37: 461–471. doi:10.1016/0041008X(76)90208-8 PMID:9709

    Kauppinen T, Toikkanen J, Pedersen D et  al. (2000).Occupational exposure to carcinogens in the European

    475

    http://dx.doi.org/10.1016/0009-2797(75)90030-7http://dx.doi.org/10.1016/0009-2797(75)90030-7http://www.ncbi.nlm.nih.gov/pubmed/1201617http://dx.doi.org/10.1016/0009-2797(77)90080-1http://dx.doi.org/10.1016/0009-2797(77)90080-1http://www.ncbi.nlm.nih.gov/pubmed/328181http://dx.doi.org/10.2307/3429293http://www.ncbi.nlm.nih.gov/pubmed/7199429http://dx.doi.org/10.1021/tx00025a001http://www.ncbi.nlm.nih.gov/pubmed/1581532http://dx.doi.org/10.1021/bi00590a023http://dx.doi.org/10.1021/bi00590a023http://www.ncbi.nlm.nih.gov/pubmed/497175http://www.ncbi.nlm.nih.gov/pubmed/7030476http://dx.doi.org/10.1016/0006-2952(79)90140-0http://www.ncbi.nlm.nih.gov/pubmed/444246http://www.ncbi.nlm.nih.gov/pubmed/8761409http://www.ncbi.nlm.nih.gov/pubmed/10626215http://dx.doi.org/10.1016/0041-008X(75)90149-0http://dx.doi.org/10.1016/0041-008X(75)90149-0http://www.ncbi.nlm.nih.gov/pubmed/813333http://dx.doi.org/10.2307/3428329http://www.ncbi.nlm.nih.gov/pubmed/1175571http://dx.doi.org/10.1021/ac60365a063http://www.ncbi.nlm.nih.gov/pubmed/1244767http://www.ncbi.nlm.nih.gov/pubmed/988738http://dx.doi.org/10.1080/15287398109530034http://www.ncbi.nlm.nih.gov/pubmed/7265317http://www.ncbi.nlm.nih.gov/pubmed/285915http://www.ncbi.nlm.nih.gov/pubmed/3482203http://www.ncbi.nlm.nih.gov/pubmed/15285078http://www.ncbi.nlm.nih.gov/pubmed/19267125http://dx.doi.org/10.1016/0006-291X(77)90599-Xhttp://www.ncbi.nlm.nih.gov/pubmed/14640http://www.ncbi.nlm.nih.gov/pubmed/6497652http://www.ncbi.nlm.nih.gov/pubmed/240126http://dx.doi.org/10.1016/0041-008X(76)90208-8http://dx.doi.org/10.1016/0041-008X(76)90208-8http://www.ncbi.nlm.nih.gov/pubmed/9709

  • IARC MONOGRAPHS – 100F

    Union. Occup Environ Med, 57: 10–18. doi:10.1136/ oem.57.1.10 PMID:10711264

    Krajewski J, Dobecki M, Gromiec J (1980). Retention ofvinyl chloride in the human lung. Br J Ind Med, 37: 373–374. PMID:7448132

    Kurliandskiĭ BA, Stovbur NN, Turusov VS (1981).Hygienic regulation of vinyl chloride Gig Sanit, 374–76. PMID:7194205

    Laib RJ & Bolt HM (1977). Alkylation of RNA by vinylchloride metabolites in vitro and in vivo: formation of 1-N(6)-etheno-adenosine. Toxicology, 8: 185–195. PMID:929626

    Langouët S, Mican AN, Müller M et  al. (1998). Misincorporation of nucleotides opposite five-membered exocyclic ring guanine derivatives byescherichia coli polymerases in vitro and in vivo:1,N2-ethenoguanine, 5,6,7,9-tetrahydro-9-oxoimidazo[1, 2-a]purine, and 5,6,7,9-tetrahydro-7-hydroxy9-oxoimidazo[1, 2-a]purine. Biochemistry, 37: 5184–5193. doi:10.1021/bi972327r PMID:9548749

    Langouët S, Müller M, Guengerich FP (1997). Misincorporation of dNTPs opposite 1,N2-ethenoguanine and 5,6,7,9-tetrahydro-7hydroxy-9-oxoimidazo[1,2-a]purine in oligonucleotides by Escherichia coli polymerases I exo- and IIexo-, T7 polymerase exo-, human immunodeficiencyvirus-1 reverse transcriptase, and rat polymerase beta.Biochemistry, 36: 6069–6079. doi:10.1021/bi962526vPMID:9166777

    Laplanche A, Clavel-Chapelon F, Contassot JC, Lanouzière CThe French VCM Group (1992). Exposureto vinyl chloride monomer: results of a cohort studyafter a seven year follow up. Br J Ind Med, 49: 134–137. PMID:1536821

    Lee CC, Bhandari JC, Winston JM et  al. (1978). Carcinogenicity of vinyl chloride and vinylidene chloride. J Toxicol Environ Health, 4: 15–30. doi:10.1080/15287397809529640 PMID:633405

    Lee FI & Harry DS (1974). Angiosarcoma of the liverin a vinyl-chloride worker. Lancet, 1(7870):1316–8.PMID:4134297 doi:10.1016/S0140-6736(74)90683-7

    Li Y, Marion MJ, Ho R et al. (2003a). Polymorphisms forvinyl chloride metabolism in French vinyl chlorideworkers. Int J Occup Med Environ Health, 16: 55–59. PMID:12705718

    Lide DR, editor (2008). CRC Handbook of Chemistry and physics. 89th ed., Boca Raton, FL: CRC Press, pp. 3–100.

    Lilly PD, Thornton-Manning JR, Gargas ML et al. (1998).Kinetic characterization of CYP2E1 inhibition in vivo and in vitro by the chloroethylenes. Arch Toxicol, 72: 609–621. doi:10.1007/s002040050551 PMID:9851676

    Maltoni C & Cotti G (1988). Carcinogenicity of vinylchloride in Sprague-Dawley rats after prenatal andpostnatal exposure. Ann N Y Acad Sci, 534: 1 Livingin a C145–159. doi:10.1111/j.1749-6632.1988.tb30108.xPMID:3389652

    Maltoni C, Lefemine G, Chieco P, Carretti D (1974). Vinylchloride carcinogenesis: current results and perspectives. Med Lav, 65: 421–444. PMID:4477887

    Maltoni C, Lefemine G, Ciliberti A et  al. (1981).Carcinogenicity bioassays of vinyl chloride monomer:a model of risk assessment on an experimental basis.Environ Health perspect, 41: 3–29. doi:10.2307/3429291 PMID:6800782

    Mastrangelo G, Fedeli U, Fadda E et al. (2004). Increasedrisk of hepatocellular carcinoma and liver cirrhosis invinyl chloride workers: synergistic effect of occupational exposure with alcohol intake. Environ Health perspect, 112: 1188–1192. PMID:15289165

    Morinello EJ, Ham AJL, Ranasinghe A et  al. (2002b). Molecular dosimetry and repair of N(2),3ethenoguanine in rats exposed to vinyl chloride.Cancer Res, 62: 5189–5195. PMID:12234983

    Morinello EJ, Koc H, Ranasinghe A, Swenberg JA (2002a).Differential induction of N(2),3-ethenoguanine in ratbrain and liver after exposure to vinyl chloride. Cancer Res, 62: 5183–5188. PMID:12234982

    Mroczkowska MM & Kuśmierek JT (1991). Miscoding potential of N2,3-ethenoguanine studied in an Escherichia coli DNA-dependent RNA polymerasein vitro system and possible role of this adduct invinyl chloride-induced mutagenesis. Mutagenesis, 6: 385–390. doi:10.1093/mutage/6.5.385 PMID:1795643

    Müller G, Norpoth K, Kusters E et al. (1978). Determinationof thiodiglycolic acid in urine specimens of vinyl chloride exposed workers. Int Arch Occup Environ Health, 41: 199–205. doi:10.1007/BF00572892 PMID:649210

    Müller M, Belas F, Ueno H, Guengerich FP (1996).Development of a mass spectrometric assay for5,6,7,9-tetrahydro-7-hydroxy-9-oximidazo[1,2-alpha]purine in DNA modified by 2-chloro-oxirane. Adv Exp Med Biol, 387: 31–36. PMID:8794191

    Mundt KA, Dell LD, Austin RP et al. (2000). Historicalcohort study of 10 109 men in the North Americanvinyl chloride industry, 1942–72: update of cancermortality to 31 December 1995. Occup Environ Med, 57: 774–781. doi:10.1136/oem.57.11.774 PMID:11024202

    Nair J, Barbin A, Guichard Y, Bartsch H (1995). 1,N6-ethenodeoxyadenosine and 3,N4-ethenodeoxycytine in liver DNA from humans and untreated rodents detected by immunoaffinity/32P-postlabeling. Carcinogenesis, 16: 613–617. PMID:7697821

    NIOSH (1990). National Occupational Exposure Survey (1981–83). Cincinnati, OH: National Institute for Occupational Safety and Health. Available at: http://www.cdc.gov/noes/noes2/76445occ.html

    NTP (2005). Vinyl ChlorideReport on Carcinogens,Eleventh Edition. Rep Carcinog, 11: 1–A32. PMID:19826456

    Osterman-Golkar S et  al. (1976). Alkylation of DNA and proteins in mice exposed to vinyl

    476

    http://dx.doi.org/10.1136/oem.57.1.10http://dx.doi.org/10.1136/oem.57.1.10http://www.ncbi.nlm.nih.gov/pubmed/10711264http://www.ncbi.nlm.nih.gov/pubmed/7448132http://www.ncbi.nlm.nih.gov/pubmed/7194205http://www.ncbi.nlm.nih.gov/pubmed/929626http://dx.doi.org/10.1021/bi972327rhttp://www.ncbi.nlm.nih.gov/pubmed/9548749http://dx.doi.org/10.1021/bi962526vhttp://www.ncbi.nlm.nih.gov/pubmed/9166777http://www.ncbi.nlm.nih.gov/pubmed/1536821http://dx.doi.org/10.1080/15287397809529640http://www.ncbi.nlm.nih.gov/pubmed/633405http://dx.doi.org/10.1016/S0140-6736(74)90683-7http://www.ncbi.nlm.nih.gov/pubmed/12705718http://dx.doi.org/10.1007/s002040050551http://www.ncbi.nlm.nih.gov/pubmed/9851676http://dx.doi.org/10.1111/j.1749-6632.1988.tb30108.xhttp://www.ncbi.nlm.nih.gov/pubmed/3389652http://www.ncbi.nlm.nih.gov/pubmed/4477887http://dx.doi.org/10.2307/3429291http://www.ncbi.nlm.nih.gov/pubmed/6800782http://www.ncbi.nlm.nih.gov/pubmed/15289165http://www.ncbi.nlm.nih.gov/pubmed/12234983http://www.ncbi.nlm.nih.gov/pubmed/12234982http://dx.doi.org/10.1093/mutage/6.5.385http://www.ncbi.nlm.nih.gov/pubmed/1795643http://dx.doi.org/10.1007/BF00572892http://www.ncbi.nlm.nih.gov/pubmed/649210http://www.ncbi.nlm.nih.gov/pubmed/8794191http://dx.doi.org/10.1136/oem.57.11.774http://www.ncbi.nlm.nih.gov/pubmed/11024202http://www.ncbi.nlm.nih.gov/pubmed/7697821http://www.ncbi.nlm.nih.gov/pubmed/19826456www.cdc.gov/noes/noes2/76445occ.html

  • Vinyl chloride

    chloride Biochemical and Biophysical Research Communications, 76 (2): 259–266. PMID:1027429doi:10.1016/0006-291X(77)90720-3

    Pirastu R, Baccini M, Biggeri A, Comba P (2003).Epidemiologic study of workers exposed to vinyl chloride in Porto Marghera: mortality update Epidemiol prev, 27: 161–172. PMID:12958735

    Pleil JD & Lindstrom AB (1997). Exhaled human breathmeasurement method for assessing exposure to halogenated volatile organic compounds. Clin Chem, 43: 723–730. PMID:9166222

    Plugge H & Safe S (1977). Vinyl chloride metabolism — a review Chemosphere, 6: 309–325. doi:10.1016/0045-6535(77)90095-9

    Radike MJ, Stemmer KL, Bingham E (1981). Effect ofethanol on vinyl chloride carcinogenesis. Environ Health perspect, 41: 59–62. doi:10.1289/ehp.814159 PMID:6277614

    Reynolds ES, Moslen MT, Szabo S, Jaeger RJ (1975). Vinylchloride-induced deactivation of cytochrome P-450and other components of the liver mixed functionoxidase system: an in vivo study. Res Commun Chem pathol pharmacol, 12: 685–694. PMID:175418

    Sabadie N, Malaveille C, Camus AM, Bartsch H (1980).Comparison of the hydroxylation of benzo(a)pyrene withthe metabolism of vinyl chloride, N-nitrosomorpholine,and N-nitroso-N’-methylpiperazine to mutagens byhuman and rat liver microsomal fractions. Cancer Res, 40: 119–126. PMID:7349891

    Simonato L, L’Abbé KA, Andersen A et  al. (1991). A collaborative study of cancer incidence and mortalityamong vinyl chloride workers. Scand J Work Environ Health, 17: 159–169. PMID:2068554

    Singer B, Kuśmierek JT, Folkman W et al. (1991). Evidencefor the mutagenic potential of the vinyl chlorideinduced adduct, N2, 3-etheno-deoxyguanosine, using asite-directed kinetic assay. Carcinogenesis, 12: 745–747. doi:10.1093/carcin/12.4.745 PMID:2013138

    Singer B, Spengler SJ, Chavez F, Kuśmierek JT (1987).The vinyl chloride-derived nucleoside, N2,3ethenoguanosine, is a highly efficient mutagen intranscription. Carcinogenesis, 8: 745–747. doi:10.1093/carcin/8.5.745 PMID:3581434

    Smulevich VB, Fedotova IV, Filatova VS (1988). Increasingevidence of the rise of cancer in workers exposed tovinylchloride. Br J Ind Med, 45: 93–97. PMID:3342200

    Summers (2006). Kirk-Othmer Encyclopedia of ChemicalTechnology. John Wiley & Sons, Inc

    Suzuki Y (1983). Neoplastic effect of vinyl chloride in mouselung–lower doses and short-term exposure. Environ Res, 32: 91–103. doi:10.1016/0013-9351(83)90195-0 PMID:6617622

    Swenberg JA, Ham A, Koc H et al. (2000). DNA adducts:effects of low exposure to ethylene oxide, vinyl chlorideand butadiene. Mutat Res, 464: 77–86. PMID:10633179

    Thériault G & Allard P (1981). Cancer mortality of a group ofCanadian workers exposed to vinyl chloride monomer.J Occup Med, 23: 671–676. doi:10.1097/00043764198110000-00009 PMID:7299502

    Thriene B, Benkwitz F, Willer H et al. (2000). Chemicalaccident in Schönbeck–an assessment of the risk to health and environment Gesundheitswesen, 62: 34–38. PMID:10705663

    Til HP, Feron VJ, Immel HR (1991). Lifetime (149-wk)oral carcinogenicity study of vinyl chloride in rats.Food Chem Toxicol, 29: 713–718. doi:10.1016/02786915(91)90130-Y PMID:1959825

    Trivers GE, Cawley HL, DeBenedetti VM et  al. (1995).Anti-p53 antibodies in sera of workers occupationally exposed to vinyl chloride. J Natl Cancer Inst, 87: 1400–1407. doi:10.1093/jnci/87.18.1400 PMID:7658501

    Ungváry G, Hudák A, Tátrai E et  al. (1978). Effects ofvinyl chloride exposure alone and in combination withtrypan blue–applied systematically during all thirdsof pregnancy on the fetuses of CFY rats. Toxicology, 11: 45–54. doi:10.1016/S0300-483X(78)90389-X PMID:705803

    Ward E, Boffetta P, Andersen A et al. (2001). Update ofthe follow-up of mortality and cancer incidence amongEuropean workers employed in the vinyl chlorideindustry. Epidemiology, 12: 710–718. PMID:11679801

    Watanabe PG, McGowan GR, Gehring PJ (1976a). Fateof (14C)vinyl chloride after single oral administration in rats. Toxicol Appl pharmacol, 36: 339–352. doi:10.1016/0041-008X(76)90013-2 PMID:1273852

    Watanabe PG, McGowan GR, Madrid EO, Gehring PJ(1976b). Fate of [14C]vinyl chloride following inhalation exposure in rats. Toxicol Appl pharmacol, 37: 49–59. doi:10.1016/S0041-008X(76)80007-5 PMID:968906

    Watanabe PG, Zempel JA, Gehring PJ (1978a). Comparisonof the fate of vinyl chloride following single and repeatedexposure in rats. Toxicol Appl pharmacol, 44: 391–399. doi:10.1016/0041-008X(78)90199-0 PMID:675709

    Watanabe PG, Zempel JA, Pegg DG, Gehring PJ (1978b).Hepatic macromolecular binding following exposureto vinyl chloride. Toxicol Appl pharmacol, 44: 571–579. doi:10.1016/0041-008X(78)90265-X PMID:684749

    Weber H, Reinl W, Greiser E (1981). German investigations on morbidity and mortality of workers exposedto vinyl chloride. Environ Health perspect, 41: 95–99. doi:10.2307/3429300 PMID:7333247

    WHO (1999). Vinyl Chloride (Environmental Health Criteria 215). Geneva: World Health Organization

    Withey JR (1976). Pharmacodynamics and uptake ofvinyl chloride monomer administered by various routes to rats. J Toxicol Environ Health, 1: 381–394. doi:10.1080/15287397609529338 PMID:1246084

    Wong O, Whorton MD, Foliart DE, Ragland D (1991).An industry-wide epidemiologic study of vinyl chloride workers, 1942–1982. Am J Ind Med, 20: 317–334. doi:10.1002/ajim.4700200305 PMID:1928109

    477

    http://dx.doi.org/10.1016/0006-291X(77)90720-3http://www.ncbi.nlm.nih.gov/pubmed/12958735http://www.ncbi.nlm.nih.gov/pubmed/9166222http://dx.doi.org/10.1016/0045-6535(77)90095-9http://dx.doi.org/10.1289/ehp.814159http://www.ncbi.nlm.nih.gov/pubmed/6277614http://www.ncbi.nlm.nih.gov/pubmed/175418http://www.ncbi.nlm.nih.gov/pubmed/7349891http://www.ncbi.nlm.nih.gov/pubmed/2068554http://dx.doi.org/10.1093/carcin/12.4.745http://www.ncbi.nlm.nih.gov/pubmed/2013138http://dx.doi.org/10.1093/carcin/8.5.745http://dx.doi.org/10.1093/carcin/8.5.745http://www.ncbi.nlm.nih.gov/pubmed/3581434http://www.ncbi.nlm.nih.gov/pubmed/3342200http://dx.doi.org/10.1016/0013-9351(83)90195-0http://www.ncbi.nlm.nih.gov/pubmed/6617622http://www.ncbi.nlm.nih.gov/pubmed/10633179http://dx.doi.org/10.1097/00043764-198110000-00009http://dx.doi.org/10.1097/00043764-198110000-00009http://www.ncbi.nlm.nih.gov/pubmed/7299502http://www.ncbi.nlm.nih.gov/pubmed/10705663http://dx.doi.org/10.1016/0278-6915(91)90130-Yhttp://dx.doi.org/10.1016/0278-6915(91)90130-Yhttp://www.ncbi.nlm.nih.gov/pubmed/1959825http://dx.doi.org/10.1093/jnci/87.18.1400http://www.ncbi.nlm.nih.gov/pubmed/7658501http://dx.doi.org/10.1016/S0300-483X(78)90389-Xhttp://www.ncbi.nlm.nih.gov/pubmed/705803http://www.ncbi.nlm.nih.gov/pubmed/11679801http://dx.doi.org/10.1016/0041-008X(76)90013-2http://www.ncbi.nlm.nih.gov/pubmed/1273852http://dx.doi.org/10.1016/S0041-008X(76)80007-5http://www.ncbi.nlm.nih.gov/pubmed/968906http://dx.doi.org/10.1016/0041-008X(78)90199-0http://www.ncbi.nlm.nih.gov/pubmed/675709http://dx.doi.org/10.1016/0041-008X(78)90265-Xhttp://www.ncbi.nlm.nih.gov/pubmed/684749http://dx.doi.org/10.2307/3429300http://www.ncbi.nlm.nih.gov/pubmed/7333247http://dx.doi.org/10.1080/15287397609529338http://www.ncbi.nlm.nih.gov/pubmed/1246084http://dx.doi.org/10.1002/ajim.4700200305http://www.ncbi.nlm.nih.gov/pubmed/1928109

  • IARC MONOGRAPHS – 100F

    Wong RH, Chen PC, Du CL et al. (2002). An increasedstandardised mortality ratio for liver cancer amongpolyvinyl chloride workers in Taiwan. Occup Environ Med, 59: 405–409. doi:10.1136/oem.59.6.405PMID:12040117

    Wong RH, Chen PC, Wang JD et al. (2003). Interactionof vinyl chloride monomer exposure and hepatitis Bviral infection on liver cancer. J Occup Environ Med, 45: 379–383. doi:10.1097/01.jom.0000063622.37065.fdPMID:12708141

    Yang Y, Nair J, Barbin A, Bartsch H (2000). Immunohistochemical detection of 1,N(6)ethenodeoxyadenosine, a promutagenic DNA adduct,in liver of rats exposed to vinyl chloride or an ironoverload. Carcinogenesis, 21: 777–781. doi:10.1093/ carcin/21.4.777 PMID:10753215

    Yllner S (1971). Metabolism of chloroacetate-1–14C inthe mouse. Basic and Clinical pharmacology and Toxicology, 30: 69–80.

    Zajdela F, Croisy A, Barbin A et al. (1980). Carcinogenicityof chloroethylene oxide, an ultimate reactive metabolite of vinyl chloride, and bis(chloromethyl)ether aftersubcutaneous administration and in initiation-promotion experiments in mice. Cancer Res, 40: 352–356. PMID:7356519

    Zhu SM, Ren XF, Wan JX, Xia ZL (2005). Evaluation invinyl chloride monomer-exposed workers and therelationship between liver lesions and gene polymorphisms of metabolic enzymes. World J Gastroenterol, 11: 5821–5827. PMID:16270392

    478

    http://dx.doi.org/10.1136/oem.59.6.405http://www.ncbi.nlm.nih.gov/pubmed/12040117http://dx.doi.org/10.1097/01.jom.0000063622.37065.fdhttp://www.ncbi.nlm.nih.gov/pubmed/12708141http://dx.doi.org/10.1093/carcin/21.4.777http://dx.doi.org/10.1093/carcin/21.4.777http://www.ncbi.nlm.nih.gov/pubmed/10753215http://www.ncbi.nlm.nih.gov/pubmed/7356519http://www.ncbi.nlm.nih.gov/pubmed/16270392

    VINYL CHLORIDE1. Exposure Data 1.1 Identification of the agent 1.2 Uses 1.3 Human exposure 1.3.1 Occupational exposure 1.3.2 Non-occupational exposure 2. Cancer in Humans 2.1 Angiosarcoma of the liver 2.2 Hepatocellular carcinoma 2.3 Cancer of the lung 2.4 Malignant neoplasms of connective and soft tissue 2.5 Other cancers 2.6 Synthesis 3. Cancer in Experimental Animals 3.1 Inhalation exposure 3.2 Oral administration 3.3 Subcutaneous and intraperitoneal injection 3.4 Transplacental administration and perinatal exposure 3.5 Carcinogenicity of metabolites 4. Other Relevant Data 4.1 Kinetics and metabolism – studies in humans 4.2 Kinetics and metabolism – studies in animals 4.3 Reaction with cellular macromolecules 4.4 Synthesis 5. Evaluation References